The Influence of Residence Time Distribution on Continuous-Flow

May 3, 2019 - Continuous-flow chemistry is emerging as an enabling technology for the synthesis of precise polymers. Recent advances in this rapidly ...
0 downloads 0 Views 2MB Size
Article Cite This: Macromolecules XXXX, XXX, XXX−XXX

pubs.acs.org/Macromolecules

The Influence of Residence Time Distribution on Continuous-Flow Polymerization Marcus H. Reis, Travis P. Varner, and Frank A. Leibfarth* Department of Chemistry, The University of North Carolina at Chapel Hill, Chapel Hill, North Carolina 27599, United States

Macromolecules Downloaded from pubs.acs.org by UNIV AUTONOMA DE COAHUILA on 05/03/19. For personal use only.

S Supporting Information *

ABSTRACT: Continuous-flow chemistry is emerging as an enabling technology for the synthesis of precise polymers. Recent advances in this rapidly growing field have hastened a need for a fundamental understanding of how fluid dynamics in tubular reactors influence polymerizations. Herein, we report a comprehensive study of how laminar flow influences polymer structure and composition. Tracer experiments coupled with in-line UV−Vis spectroscopy demonstrate how viscosity, tubing diameter, and reaction time affect the residence time distribution (RTD) of fluid in reactor geometries relevant for continuous-flow polymerizations. We found that the breadth of the RTD has strong, statistical correlations with reaction conversion, polymer molar mass, and dispersity for polymerizations conducted in continuous flow. These correlations were demonstrated to be general to a variety of different reaction conditions, monomers, and polymerization mechanisms. Additionally, these findings inspired the design of a droplet flow reactor that minimizes the RTD in continuous-flow polymerizations and enables the continuous production of well-defined polymers at a rate of 1.4 kg/day.



INTRODUCTION Continuous-flow chemistry is emerging as a useful technology for precision in polymer synthesis.1−5 Compared to conventional batch processes, the use of micromixers and small diameter tubing leads to rapid mixing, and the large specific surface area of milli- and microflow tubular reactors affords excellent heat transfer. These attributes minimize local hot spots, offer instantaneous initiation or termination of polymerizations with fast kinetics, and ultimately provide enhanced control of polymer structures and material properties.6,7 For photocatalyzed polymerizations, small diameter flow tubing allows more uniform irradiation, faster kinetics, and easier scale-up compared to batch.8−12 Lastly, software control of pumps, reactors, and in-line analytical techniques enables automation of polymerization processes.13−16 New opportunities in polymer synthesis enabled by continuous-flow chemistry include the rapid creation of polymer libraries, the telescoping of multistep polymerizations, and the automated optimization of macromolecular synthesis.17−21 In contrast to batch processes, continuous flow’s ability to dynamically alter continuous variables within a single experiment enables the rapid synthesis and analysis of polymer samples with systematic alterations in molecular weight, composition, and reaction rate. This higher-throughput approach expedites access to libraries of polymers for reaction optimization, kinetic analysis, and structure−property studies.22,23 Continuous-flow synthesis permits users to combine multistep chemical processes into an uninterrupted sequence, which has been exploited to efficiently produce multiblock and sequence-defined copolymers.18,24−28 Furthermore, the automation of continuous-flow polymerizations has the potential to © XXXX American Chemical Society

facilitate the reproducible synthesis and analysis of synthetic polymers by non-experts, analogous to peptide synthesizers, thus removing a significant barrier to material discovery efforts. The precise reaction conditions offered by continuous flow are especially relevant in controlled polymerizations where regulation of polymer microstructure, molecular weight, and dispersity (Đ) is desired. Ideally, the synergistic combination of controlled polymerizations and the well-defined reaction conditions in continuous flow leads to polymers with predictable molar mass and low Đ. Studies by Yoshida, Sawamoto, and co-workers demonstrated the advantages of continuous flow for rapid and highly exothermic cationic polymerizations where the enhanced mixing and improved heat transfer in continuous flow dramatically improved polymerization control, resulting in materials with much lower Đ than analogous reactions run in batch.29,30 For polymerizations that exhibit slower kinetics, translation of batch reactions to continuous-flow polymerization often demonstrates decreased control. In seminal work conducting reversible addition fragmentation chain-transfer (RAFT) polymerization in continuous flow, Thang and coworkers consistently observed lower monomer conversion and higher Đ compared to the analogous batch reactions for a wide variety of monomers.31 These results have been subsequently validated by a number of other independent studies and have been ascribed to fluid dynamic phenomena inherent to tubular reactors.32−36 Fully exploiting the benefits of Received: March 7, 2019 Revised: April 10, 2019

A

DOI: 10.1021/acs.macromol.9b00454 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules

Figure 1. Comparison of different flow regimes used to describe small-diameter continuous-flow tubular reactors.

plug flow, fluid velocity is assumed to be constant across any cross section of the tubular reactor, and no boundary layer is considered at the fluid−wall interface.37 The RTD of plug flow is assumed to be negligible (Figure 1). In reality, continuousflow polymerizations are conducted in tubular reactors at low to moderate flow rates under laminar flow conditions where solution in the middle of the tube exits the reactor faster than the mean residence time and solution near the walls resides in the reactor much longer. To experimentally measure the RTD of common flow geometries, in-line UV−Vis spectroscopy was used to track the dispersion of an injected tracer. A flow system was first brought to steady-state operation before a UV-absorbing small molecule tracer was introduced into the system (Figure 2A). The RTD was extracted from absorbance measurements in the time between when the tracer first begins exiting the reactor and the absorbance reaches a new steady state (Figure 2B). In a typical experiment, the observed RTD was calculated for a

continuous-flow polymer synthesis will first require a fundamental understanding of how fluid dynamics in tubular reactors influence polymer structure and composition. Modern polymerizations run in continuous-flow tubular reactors are predominately conducted in the laminar flow regime, which is defined as fluid moving through a closed channel with a Reynolds number less than 2000.37 Laminar flow is characterized by ordered, parallel layers of liquid sliding past each other at different velocities with radial mixing being limited to molecular diffusion. Due to friction at the tubing− liquid interface, a parabolic velocity profile is observed with layers at the edge moving slower than layers toward the center (Figure 1). These velocity differences are maintained over long length spans due to the lack of turbulence or recirculation eddies, causing the liquid to reside in the flow reactor for different amounts of time.38−41 The resulting residence time distribution (RTD) caused by laminar flow is defined as the distribution of time it takes simultaneously injected fluid to fully traverse and exit a flow reactor. For chemical reactions, the RTD results in a distribution in reaction time and a corresponding effect on reaction conversion and product formation.38 For polymerizations, the RTD is directly correlated to polymer Đ. Despite the emerging utility of continuous-flow polymerizations, there remains no quantitative study on how RTD influences polymer structure. Herein, we report a comprehensive study of how laminar flow in small-diameter tubular reactors influences polymer structure. Using in-line UV−Vis spectroscopy, we demonstrate how tubing diameter, solution viscosity, and mean residence time influence RTD under conditions relevant for continuousflow polymerization. We subsequently conduct polymerizations under various conditions and reactor geometries to understand how RTD influences reaction conversion, polymer molar mass (Mn), and Đ. The breadth of the RTD in laminar flow is found to have strong, statistical correlations with polymerization control. These correlations are demonstrated to be general to a variety of different reaction conditions, monomers, and polymerization mechanisms. Additionally, our comprehensive studies have enabled the rational design of a reactor to minimize RTD, enabling the continuous production of well-defined polymer at a rate of 1.4 kg/day.



RESULTS AND DISCUSSION Effect of Reactor Geometry on Experimentally Measured RTD. Reports of polymerizations conducted in continuous flow have typically assumed plug flow as a simplified model of solution moving through a reactor. In

Figure 2. (A) Continuous-flow setup used to measure residence time distribution (RTD); (B) analysis of UV−Vis response allows experimental determination of RTD.39 B

DOI: 10.1021/acs.macromol.9b00454 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules specific tubing diameter or flow rate, with hexanes as the mobile phase and toluene as the UV-absorbing small molecule tracer. To probe the effect of viscosity on RTD, the identity of the mobile phase was changed from hexanes (0.3 mPa s) to either a 90% acetonitrile−water solution (0.5 mPa s), water (1 mPa s), a 9% polyethylene glycol−water solution (9.1 mPa s), or a 23% poly(ethylene glycol)−water solution (85 mPa s). For aqueous solutions, the absorbing tracer was changed from toluene to phenol. These experiments enabled visualization of the influence of tubing diameter, residence time, and molecular diffusion on the RTD under conditions commonly used for continuous-flow polymerizations (Figure 3). In agreement with previous studies, increasing tubing size, decreasing retention time, and increasing viscosity all resulted in a broadening of RTD.38

We hypothesized that the RTD of a growing polymer would deviate from that of small molecule tracers. To test this hypothesis on a model system, we chose to explore the ringopening transesterification polymerization (ROTP) of γvalerolactone (γ-VL) catalyzed by 1,5,7-triazabicyclo[4.4.0]dec-5-ene (TBD).25 The polymerization in flow was first brought to steady-state operation using hexanol as an initiator. Subsequent replacement of hexanol with pyrenebutanol by a step change in flow rate of two pumps provided a polymerbound UV−vis tracer while keeping the total concentration of alcohol constant. The resulting RTD observed for an active polymerization (Figure 4) is much broader than those for

Figure 4. Comparison between the observed RTD of a polymerization and a small molecule model experiment under the same reaction geometry (10 min residence time and 0.04 in. tubing).

analogous experiments conducted with small molecule tracers under otherwise identical conditions. While the small-molecule experiments in Figure 3 demonstrated a symmetric RTD that is predicted by the dispersion model of fluid transport in a tubular reactor, a convective model of fluid transport more accurately describes the asymmetric RTD seen in polymerizations.39 This convective model is typically observed in flow geometries with minimal radial diffusive mixing or large tubing diameters (Figure 3A). We predict that the slower molecular diffusion of polymer chains as well as the inherent viscosity of polymer solutions leads to convective fluid transport even in small-diameter tubing, resulting in a broader RTD than previously studied small-molecule systems. Influence of RTD on Polymerization Performance. Quantifying the influence of RTD on monomer conversion, Mn, and Đ for continuous-flow polymerization is nontrivial. Viscosity constantly evolves as polymerization progresses, and different retention times have a nonlinear effect on reaction conversion. Therefore, the modification of tubing diameter was chosen as a user-defined variable to enable the systematic study of RTD’s effect on polymer structure. Model studies with small molecule tracers (Figure 3A) demonstrated that increasing tubing diameter increases the RTD; thus, we hypothesized that an increase in tubing diameter in a continuous-flow polymerization would result in a decrease in conversion, a decrease in Mn, and an increase in Đ. The controlled ROTP of γ-VL was carried out in reactors with tubing inner diameters of 0.02, 0.03, 0.04, 0.063, and 0.093 in. As a control, these results were compared to a polymerization run in a small-scale (2 mL) batch reactor, which had sufficiently fast mixing and heat transfer to approximate an experiment with no RTD. We note

Figure 3. Experimental RTD of common continuous-flow conditions using small molecule tracers. The area under the curve has been normalized to one. (A) Effect of tubing diameter on RTD with constant retention time and viscosity (10 min and 0.3 mPa s, respectively). (B) Effect of retention time on RTD with constant tubing diameter and viscosity (0.04 in. and 0.3 mPa s, respectively). (C) Effect of viscosity on RTD with constant tubing diameter and retention time (0.04 in. and 10 min, respectively). Flow rates for each condition are detailed in the Supporting Information. C

DOI: 10.1021/acs.macromol.9b00454 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules

throughput. Similarly, decreasing the tubing inner diameter (dt) exponentially increases the pressure drop and increases the difficulty of producing samples in a meaningful quantity. For these reasons, we found it challenging to use the aforementioned techniques to minimize RTD and replicate the control observed in small-scale batch polymerizations. Two alternative techniques to minimize the RTD in tubular microreactors include the creation of Dean vortices or the use of in-line static mixers.43,44 Dean vortices represent secondary fluid motion in laminar flow resulting from centripetal forces of fluid flowing in tubes arranged with tight curvature. Tracer experiments on both tightly and loosely coiled reactors showed no difference in RTD, suggesting that Dean vortices did not occur in the geometries and reaction conditions explored herein. The use of static mixers is a potential solution, but they exaggerate ΔP, are prone to clog with viscous solutions, and require optimization for each system. Gas−liquid segmented flow, also referred to as Taylor flow or droplet flow,45 is an alternative and attractive approach to decrease the RTD in continuous-flow reactors. The droplet flow approach compartmentalizes slugs of liquid in an immiscible mobile phase (Figure 5). Within each slug,

that even moderate scale-up of batch reactions considerably diminishes mass and heat transfer and leads to loss of control,42 but a comprehensive study of these effects is outside the scope of this manuscript. As shown in Table 1, the RTD in laminar flow led to decreased control compared to an analogous small-scale batch Table 1. Continuous-Flow Polymerization of γ-VL with Varied Tubing Diametersa

reactor geometryb batch (2 mL scale) flow (0.02 in. tubing) flow (0.03 in. tubing) flow (0.04 in. tubing) flow (0.063 in. tubing) flow (0.093 in. tubing)

11,200 11,300 11,000 10,800 9000 9000

± ± ± ± ± ±

200 900 400 500 300 1000

conversion (%)d

Đc

Mn (g/mol)c 1.05 1.12 1.14 1.15 1.27 1.33

± ± ± ± ± ±

0.01 0.01 0.04 0.01 0.07 0.06

69 69 63 62 50 49

± ± ± ± ± ±

9 2 3 2 5 8

a

Conditions: [γ-VL] = 5 M; [γ-VL]/[initiator]/[TBD] = 125:1:0.4; theoretical Mn at 100% conversion = 12,500 g/mol. Reactions were conducted in triplicate, and standard deviations are reported. bTubing lengths and flow rates found in the Supporting Information. c Measured by GPC using polystyrene standards. dDetermined by 1 H NMR.

reaction, with clear correlations observed between tubing diameter, conversion, Mn, and Đ. Each experiment was conducted three times and standard deviations are reported in Table 1. Even in the smallest inner-diameter tubing tested, Đ increased from 1.05 in batch to 1.12 in flow, and increases in tubing inner-diameter led to further increases in Đ. In the largest diameter tubing (0.093 in.), both conversion and Mn decreased 20% compared to batch and Đ increased from 1.05 to 1.33. These results represent a significant limitation for the scale-up of continuous-flow polymerizations, since large tubing inner-diameters enable higher continuous throughput for a given reaction. Overcoming RTD with Droplet Flow. The combined UV−Vis and polymerization data suggest that decreasing tubing diameter, increasing retention time, and decreasing viscosity result in better control. For polymerizations run in continuous flow, however, the pressure drop along a given length of tubing must be considered concomitantly with changes to the reactor geometry. A pressure drop (ΔP) is the result of frictional forces between the stationary tubing and the moving liquid and can be quantified by the Hagen−Poiseuille equation37 ΔP =

Figure 5. Polymerization of γ-VL using droplet flow. The reaction run in 0.093 in. inner diameter tubing produced poly(γ-VL) at a rate of 1.4 kg/day. Ar = argon. Conditions: [γ-VL] = 5 M; [γ-VL]/ [initiator]/[TBD] = 125:1:0.4; theoretical Mn at 100% conversion = 12,500 g/mol. aTubing lengths and flow rates found in the Supporting Information. bMeasured by GPC using polystyrene standards. cDetermined by 1H NMR.

recirculation motions enhance mixing, and the lack of slug coalescence ensures a narrow RTD. This approach has been shown to narrow the RTD of both small molecule reactions and nanoparticle syntheses; for the latter, the size dispersity of nanoparticles formed in droplet flow decreased compared to continuous flow.46−48 Using a metering valve and an inert gas, we evaluated the influence of droplet flow on the RTD of continuous-flow polymerizations. Tracer experiments analyzed by in-line UV− Vis spectroscopy were used to quantify the RTD of viscous fluids in droplet flow. For the most viscous sample analyzed in Figure 3C (85 mPa s), droplet flow resulted in a narrower RTD than even the analogous nonviscous sample (0.3 mPa s) run in traditional continuous flow. The polymerization of γ-VL was used to probe the influence of droplet flow on polymerization (Figure 5). Droplet flow demonstrated a dramatic improvement compared to analogous

128μLQ πd t 4

where μ is the dynamic viscosity, L is the tubing length, Q is the volumetric flow rate, and dt is the tubing diameter. For polymerizations conducted in continuous flow, a pressure drop commonly becomes high enough to stall pumps or burst tubing due the increasing viscosity of the polymerizing mobile phase. Strategies to mitigate the negative impacts of laminar flow on polymer structure and Đ also exaggerate the pressure drop within the flowing system. For example, diluting the reaction will lower the pressure drop but requires a faster flow rate (Q) and longer lengths of tubing (L) to maintain an equal D

DOI: 10.1021/acs.macromol.9b00454 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules

an even lower Đ than that observed in an analogous batch experiment (Table S2). The polymerization of DMA, however, demonstrated additional complexities of running polymerizations in batch or continuous flow for very rapid polymerizations (Table S3). For the polymerization of DMA, heat transfer and mixing were found to have significantly larger effects on conversion and Đ than RTD. For example, in two otherwise identical reactions, conversion in a 2 mL batch reactor was increased from 6 to 87% upon preheating the monomer solution prior to adding the thermal initiator, highlighting the slow heat transfer even in small-scale batch reactions. The analogous flow reaction achieved a moderate conversion of 64% without any preheating. Surprisingly, attempts to preheat the flow reaction by separating the monomer and initiator solutions into two syringes and recombining them using a T-mixer led to a decrease in conversion and an increase in Đ. The collection of these experimental observations suggests that mixing, and not the reactor geometry, is the primary determinant to achieve controlled polymerizations in flow with monomers that exhibit rapid kinetics. For fast reactions in flow, improving mixing efficiency through the use of more advanced micromixers is suggested, and a number of quality discussions on the topic can be found elsewhere.38,39

continuous-flow polymerizations. Even in the largest diameter tubing, the conversion, Mn, and Đ all effectively matched that of a small-scale batch reaction. The improvement in RTD was found to be decoupled from retention time, viscosity, and tubing diameter, thus overcoming a crucial hurdle for continuous-flow polymerizations. Using two peristaltic pumps at a total flow rate of 3.6 mL/s and 0.093 in. inner-diameter tubing, the polymerization of γ-VL was scaled up to collect poly(γ-VL) continuously at a rate that would equate to 1.4 kg/ day under continuous operation. Importantly, both the Mn (12.0 kg/mol) and Đ (1.07) of the polymer were analogous to those of a small-scale batch polymerization. A built-in pressure monitoring system demonstrated that the total system pressure never exceeded 5 psi, indicating that further scale-up is possible. In order to probe the generality of RTD’s effects, we studied the RAFT polymerization of n-butyl acrylate (nBuA), Nisopropylacrylamide (NIPAM), and N,N-dimethylacrylamide (DMA). These studies expanded the data set from ROTP to controlled radical polymerization while simultaneously investigating polymerizations with both slower (nBuA) and faster (NIPAM and DMA) reaction kinetics. The RAFT polymerization of nBuA was studied at a retention time of 30 min (Table 2). The polymerization run in batch resulted in



Table 2. Continuous-Flow Polymerization of nBuA with Varied Tubing Diametersa

reactor geometryb

Mn (g/mol)c

Đc

conversion(%)d

batch (2 mL scale) flow (0.04 in. tubing) flow (0.093 in. tubing) droplet (0.093 in tubing)

13,400 13,100 13,100 13,600

1.14 1.17 1.26 1.15

62 64 65 60

CONCLUSIONS We have provided experimental evidence that the RTD associated with laminar flow in tubular microreactors has strong, statistical correlations with reaction conversion, Mn, and Đ in continuous-flow polymerizations. Tracer experiments coupled with in-line UV−Vis spectroscopy demonstrated that decreasing retention time, increasing viscosity, and increasing tubing diameter resulted in broadening of the RTD. For the ROTP of γ-VL, we found that the broadening of the RTD directly correlated to an increase in Đ and a decrease in both Mn and conversion. Droplet flow was discovered as a solution to mitigate the negative effects of RTD for the continuous-flow polymerization of γ-VL, resulting in a method that matched the results of a small-scale batch reaction while producing narrowly dispersed polymer at a rate of 1.4 kg/day. Furthermore, these observations were generalized to polymerizations with different mechanisms and different reaction rates. The fundamental understanding of laminar flow’s influence on polymerization detailed herein will facilitate the broader adoption of continuous-flow synthesis in polymer science.

a

Conditions: [nBuA] = 2 M; [nBuA]/[CTA]/[AIBN] = 200:1:0.3; theoretical Mn at 100% conversion = 25,600 g/mol. bTubing lengths and flow rates found in the Supporting Information. cMeasured by GPC using polystyrene standards. dDetermined by 1H NMR.

poly(nBuA) with an Mn of 13.0 kg/mol and Đ of 1.14. Polymerizations run in continuous flow demonstrated a correlation between increasing tubing diameter and increasing Đ. This loss of polymerization control is consistent with the results observed with poly(γ-VL) (Table 1). We hypothesize that the magnitude of the change in Đ is less for the slower polymerization of nBuA due to the narrower RTD observed for reactions with longer residence times (Figure 3B). The polymerization of nBuA, however, did not demonstrate a reduction in monomer conversion with increasing tubing diameter. We hypothesize that the broadening of Đ without the associated decrease in monomer conversion is due to the differences in the RTD of the monomers and growing polymer chains in solution. Due to differences in molecular diffusion, a narrower RTD is expected for unreacted monomers as compared to the growing polymer chains, especially at longer residence times (Figure. 4). This would lead to the RTD in this system having a smaller impact on monomer conversion as compared to polymer Đ. Expected trends were observed for the polymerization of NIPAM at 5 min retention time, with droplet flow leading to



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.macromol.9b00454. Detailed information on reagents, analytic instrumentation, specific reactor tubing specifications, flow rates, list of commercially available flow equipment, specific procedures for the polymerizations of δ-valerolactone, n-butyl acrylate, N-isopropylacrylamide, and N,N-dimethylacrylamide, extended polymerization data of poly(nbutyl acrylate), poly(N-isopropylacrylamide), poly(N,Ndimethylacrylamide), and explanation of data workup for residence time distribution visualizations (PDF) E

DOI: 10.1021/acs.macromol.9b00454 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules



Continuous Flow Aminolysis of RAFT Polymers Using Multistep Processing and Inline Analysis. Macromolecules 2014, 47, 8203−8213. (18) Corrigan, N.; Manahan, R.; Lew, Z. T.; Yeow, J.; Xu, J.; Boyer, C. Copolymers with Controlled Molecular Weight Distributions and Compositional Gradients through Flow Polymerization. Macromolecules 2018, 51, 4553−4563. (19) Rosenfeld, C.; Serra, C.; O’Donohue, S.; Hadziioannou, G. Continuous Online Rapid Size Exclusion Chromatography Monitoring of Polymerizations - CORSEMP. Macromol. React. Eng. 2007, 1, 547−552. (20) Reis, M. H.; Davidson, C. L. G., IV; Leibfarth, F. A. Continuous-Flow Chemistry for the Determination of Comonomer Reactivity Ratios. Polym. Chem. 2018, 9, 1728−1734. (21) Corrigan, N.; Almasri, A.; Taillades, W.; Xu, J.; Boyer, C. Controlling Molecular Weight Distributions through Photoinduced Flow Polymerization. Macromolecules 2017, 50, 8438−8448. (22) Brocken, L.; Price, P. D.; Whittaker, J.; Baxendale, I. R. Continuous Flow Synthesis of Poly(Acrylic Acid) via Free Radical Polymerisation. React. Chem. Eng. 2017, 2, 662−668. (23) Krishnadasan, S.; Brown, R. J. C.; Demello, A. J.; Demello, J. C. Intelligent Routes to the Controlled Synthesis of Nanoparticles. Lab Chip 2007, 7, 1434−1441. (24) Kuroki, A.; Martinez-Botella, I.; Hornung, C. H.; Martin, L.; Williams, E. G. L.; Locock, K. E. S.; Hartlieb, M.; Perrier, S. Looped Flow RAFT Polymerization for Multiblock Copolymer Synthesis. Polym. Chem. 2017, 8, 3249−3254. (25) Zhu, N.; Feng, W.; Hu, X.; Zhang, Z.; Fang, Z.; Zhang, K.; Li, Z.; Guo, K. Organocatalyzed Continuous Flow Ring-Opening Polymerizations to Homo- and Block-Polylactones. Polymer 2016, 84, 391−397. (26) Wicker, A. C.; Leibfarth, F. A.; Jamison, T. F. Flow-IEG Enables Programmable Thermodynamic Properties in SequenceDefined Unimolecular Macromolecules. Polym. Chem. 2017, 8, 5786. (27) Baeten, E.; Haven, J. J.; Junkers, T. RAFT Multiblock Reactor Telescoping: From Monomers to Tetrablock Copolymers in a Continuous Multistage Reactor Cascade. Polym. Chem. 2017, 8, 3815−3824. (28) Van De Walle, M.; De Bruycker, K.; Junkers, T.; Blinco, J. P.; Barner-Kowollik, C. Scalable Synthesis of Sequence-Defined Oligomers via Photoflow Chemistry. ChemPhotoChem 2019, DOI: 10.1002/cptc.201900031. (29) Nagaki, A.; Iwasaki, T.; Kawamura, K.; Yamada, D.; Suga, S.; Ando, T.; Sawamoto, M.; Yoshida, J.-i. Microflow System Controlled Carbocationic Polymerization of Vinyl Ethers. Chem. - Asian J. 2008, 3, 1558−1567. (30) Nagaki, A.; Kawamura, K.; Suga, S.; Ando, T.; Sawamoto, M.; Yoshida, J.-i. Cation Pool-Initiated Controlled/Living Polymerization Using Microsystems. J. Am. Chem. Soc. 2004, 126, 14702−14703. (31) Hornung, C. H.; Guerrero-Sanchez, C.; Brasholz, M.; Saubern, S.; Chiefari, J.; Moad, G.; Rizzardo, E.; Thang, S. H. Controlled RAFT Polymerization in a Continuous Flow Microreactor. Org. Process Res. Dev. 2011, 15, 593−601. (32) Paulus, R. M.; Erdmenger, T.; Becer, C. R.; Hoogenboom, R.; Schubert, U. S. Scale-Up of Microwave-Assisted Polymerizations in Continuous-Flow Mode: Cationic Ring-Opening Polymerization of 2Ethyl-2-Oxazoline. Macromol. Rapid Commun. 2007, 28, 484−491. (33) Russum, J. P.; Jones, C. W.; Schork, F. J. Continuous Reversible Addition-Fragmentation Chain Transfer Polymerization in Miniemulsion Utilizing a Multi-Tube Reaction System. Macromol. Rapid Commun. 2004, 25, 1064−1068. (34) Wurm, F.; Wilms, D.; Klos, J.; Löwe, H.; Frey, H. Carbanions on Tap − Living Anionic Polymerization in a Microstructured Reactor. Macromol. Chem. Phys. 2008, 209, 1106−1114. (35) Iida, K.; Chastek, T. Q.; Beers, K. L.; Cavicchi, K. A.; Chun, J.; Fasolka, M. J. Living Anionic Polymerization Using a Microfluidic Reactor. Lab Chip 2009, 9, 339−345. (36) Diehl, C.; Laurino, P.; Azzouz, N.; Seeberger, P. H. Accelerated Continuous Flow RAFT Polymerization. Macromolecules 2010, 43, 10311−10314.

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Frank A. Leibfarth: 0000-0001-7737-0331 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS Financial support for this work was provided by the 3M Nontenured Faculty Award and UNC startup funds.



REFERENCES

(1) Tonhauser, C.; Natalello, A.; Löwe, H.; Frey, H. Microflow Technology in Polymer Synthesis. Macromolecules 2012, 45, 9551− 9570. (2) Myers, R. M.; Fitzpatrick, D. E.; Turner, R. M.; Ley, S. V. Flow Chemistry Meets Advanced Functional Materials. Chem. - Eur. J. 2014, 12348−12366. (3) Plutschack, M. B.; Pieber, B.; Gilmore, K.; Seeberger, P. H. The Hitchhiker’s Guide to Flow Chemistry. Chem. Rev. 2017, 117, 11796−11893. (4) Junkers, T. Precise Macromolecular Engineering via ContinuousFlow Synthesis Techniques. J. Flow Chem. 2017, 7, 106−110. (5) Hu, X.; Zhu, N.; Fang, Z.; Guo, K. Continuous Flow RingOpening Polymerizations. React. Chem. Eng. 2017, 2, 20−26. (6) Junkers, T. Precision Polymer Design in Microstructured Flow Reactors: Improved Control and First Upscale at Once. Macromol. Chem. Phys. 2017, 218, 1600421. (7) Iwasaki, T.; Yoshida, J.-i. Free Radical Polymerization in Microreactors. Significant Improvement in Molecular Weight Distribution Control. Macromolecules 2005, 38, 1159−1163. (8) Chen, M.; Zhong, M.; Johnson, J. A. Light-Controlled Radical Polymerization: Mechanisms, Methods, and Applications. Chem. Rev. 2016, 116, 10167−10211. (9) Melker, A.; Fors, B. P.; Hawker, C. J.; Poelma, J. E. Continuous Flow Synthesis of Poly(Methyl Methacrylate) via a Light-Mediated Controlled Radical Polymerization. J. Polym. Sci., Part A: Polym. Chem. 2015, 53, 2693−2698. (10) Wenn, B.; Conradi, M.; Carreiras, A. D.; Haddleton, D. M.; Junkers, T. Photo-Induced Copper-Mediated Polymerization of Methyl Acrylate in Continuous Flow Reactors. Polym. Chem. 2014, 5, 3053−3060. (11) Corrigan, N.; Rosli, D.; Jones, J. W. J.; Xu, J.; Boyer, C. Oxygen Tolerance in Living Radical Polymerization: Investigation of Mechanism and Implementation in Continuous Flow Polymerization. Macromolecules 2016, 49, 6779−6789. (12) Ramsey, B. L.; Pearson, R. M.; Beck, L. R.; Miyake, G. M. Photoinduced Organocatalyzed Atom Transfer Radical Polymerization Using Continuous Flow. Macromolecules 2017, 50, 2668− 2674. (13) Rubens, M.; Vrijsen, J. H.; Laun, J.; Junkers, T. Precise Polymer Synthesis by Autonomous Self-Optimizing Flow Reactors. Angew. Chem., Int. Ed. 2019, 58, 3183−3187. (14) Bally, F.; Serra, C. A.; Brochon, C.; Anton, N.; Vandamme, T.; Hadziioannou, G. A Continuous-Flow Polymerization Microprocess with Online GPC and Inline Polymer Recovery by MicromixerAssisted Nanoprecipitation. Macromol. React. Eng. 2011, 5, 542−547. (15) Bédard, A.-C.; Adamo, A.; Aroh, K. C.; Russell, M. G.; Bedermann, A. A.; Torosian, J.; Yue, B.; Jensen, K. F.; Jamison, T. F. Reconfigurable System for Automated Optimization of Diverse Chemical Reactions. Science 2018, 361, 1220−1225. (16) Haven, J. J.; Junkers, T. Online Monitoring of Polymerizations: Current Status. Eur. J. Org. Chem. 2017, 6474−6482. (17) Hornung, C. H.; von Känel, K.; Martinez-Botella, I.; Espiritu, M.; Nguyen, X.; Postma, A.; Saubern, S.; Chiefari, J.; Thang, S. H. F

DOI: 10.1021/acs.macromol.9b00454 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules (37) Spurk, J. H.; Aksel, N. Fluid Mechanics; Second ed. Springer: VerlagBerlinHeidelberg, 2008. (38) Nagy, K. D.; Shen, B.; Jamison, T. F.; Jensen, K. F. Mixing and Dispersion in Small-Scale Flow Systems. Org. Process Res. Dev. 2012, 16, 976−981. (39) Gobert, S. R. L.; Kuhn, S.; Braeken, L.; Thomassen, L. C. J. Characterization of Milli- and Microflow Reactors: Mixing Efficiency and Residence Time Distribution. Org. Process Res. Dev. 2017, 21, 531−542. (40) Levenspiel, O.; Smith, W. K. Notes on the Diffusion-Type Model for the Longitudinal Mixing of Fluids in Flow. Chem. Eng. Sci. 1957, 6, 227−235. (41) Taylor, G. I. Dispersion of Soluble Matter in Solvent Flowing Slowly through a Tube. Proc. R. Soc. London, Ser. A 1953, 219, 186− 203. (42) Meyer, T. Scale-up of Polymerization Process: A Practical Example. Org. Process Res. Dev. 2003, 7, 297−302. (43) Nivedita, N.; Ligrani, P.; Papautsky, I. Dean Flow Dynamics in Low-Aspect Ratio Spiral Microchannels. Sci. Rep. 2017, 7, 44072. (44) Lueth, F. G.; Pauer, W.; Moritz, H.-U. Properties of SmartScaled PTFE-Tubular Reactors for Continuous Emulsion Polymerization Reactions. Macromol. Symp. 2013, 333, 69−79. (45) Song, H.; Chen, D. L.; Ismagilov, R. F. Reactions in Droplets in Microfluidic Channels. Angew. Chem., Int. Ed. 2006, 45, 7336−7356. (46) Sebastian Cabeza, V.; Kuhn, S.; Kulkarni, A. A.; Jensen, K. F. Size-Controlled Flow Synthesis of Gold Nanoparticles Using a Segmented Flow Microfluidic Platform. Langmuir 2012, 28, 7007− 7013. (47) Yen, B. K. H.; Günther, A.; Schmidt, M. A.; Jensen, K. F.; Bawendi, M. G. A Microfabricated Gas-Liquid Segmented Flow Reactor for High-Temperature Synthesis: The Case of CdSe Quantum Dots. Angew. Chem., Int. Ed. 2005, 44, 5447−5451. (48) Köhler, J. M.; Li, S.; Knauer, A. Why Is Micro Segmented Flow Particularly Promising for the Synthesis of Nanomaterials? Chem. Eng. Technol. 2013, 36, 887−899.

G

DOI: 10.1021/acs.macromol.9b00454 Macromolecules XXXX, XXX, XXX−XXX